through Dispersion-Corrected DFT and NMR ... - ACS Publications

Jul 14, 2014 - Some revisions to the original assignment of the 13C/1H NMR δiso parameters of 1·H2O, 1β, and 1γ are ... Journal of Chemical Theory...
1 downloads 0 Views 3MB Size
Article pubs.acs.org/IC

Unraveling the Polymorphism of [(p‑cymene)Ru(κN-INA)Cl2] through Dispersion-Corrected DFT and NMR GIPAW Calculations Davide Presti, Alfonso Pedone,* and Maria Cristina Menziani* Dipartimento di Scienze Chimiche e Geologiche, Università di Modena e Reggio-Emilia, 183 via G. Campi, I-41125 Modena, Italy S Supporting Information *

ABSTRACT: The structural and 13C/1H NMR parameters of the four crystal forms (1α, 1·H2O, 1β, and 1γ) of the solid wheel-and-axle (WAA) metal−organic compound [(p-cymene)Ru(κN-INA)Cl2] have been studied by means of periodic DFT calculations. The quality of the results obtained strongly depends on a correct description of long-range interactions; thus, in the geometry refinement protocol used, the pure DFT functionals need to be coupled with a dispersion-correction term (B3LYPD2, B3LYP-D*). The solid-state 13C/1H NMR δiso parameters and 13C MAS NMR spectra, calculated by means of the PBEGIPAW method, agree well with the experimental data for the four crystal forms (mean absolute deviations of the 13C and 1H δiso data values lie in the ranges 1.3−2.9 and 0.3−1.0 ppm, respectively). In this context, some revisions in the experimental assignment of the 13C/1H NMR δiso parameters of the 1·H2O, 1β, and 1γ crystal forms can be suggested. The mismatch in the assignment seems to be due to the rotation of the −COOH moiety, which occurs at the 1α−1·H2O transition and was not considered in the experiments. Finally, the results obtained suggest the presence of two COOH···Cl hydrogen bonds of comparable strength established by the two molecules in the asymmetric unit of the 1γ polymorph, in partial disagreement with previous findings.



INTRODUCTION Wheel-and-axle (WAA) compounds represent a class of organic/inorganic molecules characterized by a long, usually linear spacer (axle) and two bulky groups (wheels) at both ends. The shape and the irregularity of these types of structures promote the generation of a widespread supramolecular organization governed by hydrogen bonds and/or dispersive interactions. Therefore, WAAs present well-known host−guest properties and a high tendency to form clathrates with different solvents,1 thus attracting the interest of many researchers to the design of WAA compounds for applications in heterogeneous catalysis, solid-state “green” chemistry, sensoring, and storage of gases.2−5 Very recently, Bacchi and co-workers2 reported the design of a new series of WAAMOs based on the half-sandwich Ru(II) complex [(p-cymene)Ru(κN-INA)Cl2] (INA = isonicotinic acid) with the aim of obtaining a supramolecular synthon by cyclic supramolecular dimerization of the INA carboxylic groups. The solid [(p-cymene)Ru(κN-INA)Cl2] was obtained in four different crystalline forms: 1α, 1·H2O, 1β, and 1γ. The isolation of the pure metastable 1γ polymorph and its XRPD/NMR characterization were achieved in a subsequent work by the same authors6 who, in addition, realized that the 1β form is also metastable and, if cooled down slowly, transforms into 1γ, not 1α as observed previously. Moreover, some peculiarities in the structure of the 1β polymorph were also reconsidered.6 © 2014 American Chemical Society

Interestingly, the X-ray single-crystal and powder diffraction, ATR-FTIR, and solid-state NMR (ssNMR) spectroscopic characterization revealed that the main intermolecular interactions governing the supramolecular architecture of 1α, 1β, and 1γ was not the expected cyclic supramolecular dimer involving the −COOH functions but rather the hydrogen bonds between the chloride ligands (acceptor) bonded to Ru(II) and the carboxylic OH moiety (donor). Moreover, for the 1·H2O form the cyclic supramolecular dimerization of the COOH groups of INA is prevented by the inclusion of water. All of these observations give an idea of the complexity of the [(p-cymene)Ru(κN-INA)Cl2] polymorphism, knowledge of which represents an important challenge, since it implies the characterization of the weak noncovalent interactions determining one molecular packing with respect to the others. NMR crystallography, that is, the combination of solid-state NMR spectroscopic data with X-ray powder diffraction (XRPD) methods, is a powerful tool to investigate the structure of materials difficult to obtain as large single crystals.7 However, the assignment of the observed 13C and 1H solidstate NMR resonances to the atoms belonging to distinct molecules in the asymmetric unit cell is a considerable challenge that can only be met by the help of computational NMR. The successful prediction of chemical shifts and Received: March 23, 2014 Published: July 14, 2014 7926

dx.doi.org/10.1021/ic5006743 | Inorg. Chem. 2014, 53, 7926−7935

Inorganic Chemistry

Article

thresholds for Coulomb and exchange series accuracy (keyword TOLINTEG; see ref 35) were set to 10−8, 10−8, 10−8, 10−10, and 10−25. NMR Calculations. Calculations of NMR parameters were carried out on the crystal structures previously optimized at the B3LYP, B3LYP-D2, and B3LYP-D* level of theory. The 1H and 13C shielding tensors were computed by means of the gauge including projection augmented wave (GIPAW) approach44,45 encoded in the CASTEP software46,47 which uses plane waves as the basis set and the pseudopotential approach to describe the core− valence interactions. All calculations were performed using the PBE exchange-correlation functional48 without including dispersion-correction terms. In fact, being a posteriori empirical corrections to the DFT energy and forces, they do not directly affect the computation of the NMR parameters. Ultrasoft pseudopotentials included in the CASTEP library were used, and the plane wave basis set was cut at 610 eV. A Monkhorst−Pack grid of minimum sample spacing of 0.04 Å−1 in the reciprocal lattice was adopted. In order to compare the isotropic shielding directly with experimentally measured isotropic chemical shifts, the following relations were used for 13C and 1H, respectively:

quadrupolar coupling constants of various inorganic crystalline and amorphous solids8−14 as well as of organic molecular crystals and biomolecules8,15,16 has been achieved by means of recent developments of the gauge including projector augmented wave (GIPAW) method. GIPAW is especially devised for the computation of NMR parameters in extended solids described within periodic boundary conditions, plane wave basis sets, and density functional theory (DFTGIPAW).17,18 The accuracy of the NMR calculation results is strongly dependent on the quality of the structural coordinates employed.19 Cost-effective energy minimization approaches based on density functional theory can be employed to generate refined structures20−23 to achieve the best agreement between experimentally measured and ab initio calculated chemical shift tensor components. However, the most common exchangecorrelation functionals, based on semilocal electron correlation, do not accurately describe the effects of the dispersive forces, which govern the packing arrangements and the stability of molecular solids. Thus, London-type pairwise corrections based on either empirical24,25 or nonempirical26−30 parameters or nonlocal DFT functionals31−33 must be applied.34,35 In this paper, the solid-state 13C/1H NMR δiso parameters and 13C MAS NMR spectra of the 1α, 1·H2O, 1β, and 1γ crystal forms of [(p-cymene)Ru(κN-INA)Cl2] will be calculated by means of the PBE-GIPAW method and the results will be compared with the experimental data. The importance of the quality of the structural coordinates used (i.e., the correct treatment of dispersion interactions in the refinement of the crystal structure of the crystal forms) will be discussed.



δiso(13C) = − 0.960σiso + 166.6 δiso(1H) = − 0.944σiso + 29.6 which were fitted on the measured isotropic chemical shift of the 1α crystal phase that was taken as the reference. The linear regression shown in Figure S1 of the Supporting Information yielded R2 values of 0.999 and 0.978 for 13C and 1H, respectively. The computed 13C isotropic chemical shifts were then employed to simulate the corresponding MAS NMR spectra of all the crystal forms, by means of a homemade code based on the resolution of spin effective Hamiltonians.49 The 1H MAS NMR spectra were not simulated because the broadening effects due to the homonuclear dipole−dipole interactions, which are not averaged out in the experimental spectra reported in the literature,6 are not included in the present version of our code.

COMPUTATIONAL PARAMETERS



Geometry Optimizations. Gaussian-basis DFT calculations were carried out by using a parallel version of the CRYSTAL09 package,36,37 imposing periodic boundary conditions (PBC). The starting geometries of the 1α and 1·H2O crystal forms were taken from ref 2, whereas those of the 1β and 1γ crystal forms correspond to the nonrefined geometries of ref 6. Full geometry optimizations (lattice parameters and atomic displacements) of the 1α, 1·H2O, 1β, and 1γ crystal forms were performed by using the hybrid B3LYP functional,38,39 its combination with the “classical” Grimme dispersion correction (B3LYP-D2),24 and the alternative reparametrization of the Grimme D2 term devised for crystals (B3LYP-D*).40 Moreover, for the 1β and 1γ crystal forms B3LYP-D* geometry optimizations with allowance for the relaxation of the positions of the hydrogen atoms only were also carried out. The basis set chosen furnishes an optimal characterization of the electronic configuration of the Ru(II) metal−organic synthons and their arrangement within a manifold crystalline framework. Chlorine and ruthenium atoms were described by modified fully relativistic pseudopotentials on core electrons (MDF28 ECP for Ru,41 SDF10 ECP for Cl42primitive Gaussians with exponents lower than 0.06 have been excluded to avoid basis set linear dependence), and associated double-ζ quality basis sets on valence electrons, taken from the Stuttgart pseudopotential library.43 The 6-31G(d) basis set was used for H, C, N, and O atoms; a ptype polarization was added on H atoms and sp-type diffuse functions (consistent with the 6-31+G(d) basis set) were used for N and O atoms. The use of diffuse functions is justified by the fact that N atoms are directly involved in the Ru(II) complex, whereas O atoms can form, depending on the case, the hydrogen bond. The Monkhorst−Pack grid of k points was generated with a shrinking factor of 2 (keyword SHRINK 2 2), that corresponds to 8 k points inside the Irreducible Brillouin Zone (IBZ), whereas the DFT integration grid consisted of 75 radial and 974 angular points. The

RESULTS AND DISCUSSION 1α and 1·H2O Crystal Forms. Fully Optimized Structures. The 1α phase belongs to a monoclinic system (P21/n), it contains four molecules of [(p-cymene)Ru(κN-INA)Cl2] per unit cell (164 atoms), and it is characterized by intermolecular interactions represented by the −COOH···Cl− hydrogen bond along the 21 axis, which involve two consecutive molecules arranged in two alternate zigzag chains (a side view is displayed in Figure 1). The inclusion of water (four molecules per unit cell) has a strong structural role in the packing of 1·H2O, since it organizes the half-sandwich units in water-assisted centrosymmetric supramolecular dimers. The water molecules bridge pairs of carboxylic groups by the OH···O and OH···Cl interactions: the metal−arene wheels interact together, forming the “inverted piano-stool” supramolecular motif,2 which is clearly shown in Figure S2 of the Supporting Information. Cell, Intra- and Intermolecular Parameters. An extensive analysis of the full optimized geometry of the 1α and 1·H2O crystal forms, performed by using the hybrid B3LYP functional,38,39 its combination with the “classical” Grimme dispersion correction (B3LYP-D2),24 and the alternative reparametrization of the Grimme D2 term devised for crystals (B3LYP-D*)40 has been carried out; the results are reported in detail in Tables S1 and S2 and in Figure S3 of the Supporting Information. Since the B3LYP-D* functional provides the best performance in reproducing the experimental structure, for brevity, only the results obtained by the B3LYP-D* full optimization are given in Tables 1 and 2. 7927

dx.doi.org/10.1021/ic5006743 | Inorg. Chem. 2014, 53, 7926−7935

Inorganic Chemistry

Article

Table 2. Cell Parameters and Important Angles and Distances of the 1·H2O Form Obtained from Full Geometry Optimizationsa a (Å) b (Å) c (Å) β (deg) V (Å3) H2O···(H)O (Å)c HO(H)···O (Å)d HO(H)···Cl (Å)e N−Ru−p-cymene (deg) Ru−p-cymene (Å) Cl−Ru−Cl (deg) INA ring parallel (Å) INA ring alternate (Å)

B3LYP-D*

exptlb

15.146 (−3.684) 15.753 (−0.544) 7.257 (−2.204) 87.830 (4.280) 1730.14 (−5.908) 2.542 (0.014) 2.781 (0.027) 3.103 (0.007) 129.061 (0.115) 1.685 (0.017) 86.873 (0.856) 7.257 (0.163) 3.880 (0.089)

15.725 15.839 7.420 84.225 1838.78 2.556 2.808 3.110 128.946 1.668 86.017 7.420 3.969

a

Percent relative deviations are reported in parentheses for the cell parameters, and absolute deviations are reported in parentheses for the distances and angles. bFrom ref 2. cH bond between water oxygen and −COOH oxygen (hydrogen). dH bond between water oxygen (hydrogen) and −COOH oxygen. eH bond between water oxygen (hydrogen) and chlorine.

Figure 1. Perspective view of the 1α polymorph (B3LYP-D*), with the continuous black line representing the unit cell. The Cl···O distance and the C−OCl angle are displayed, and hydrogen bonds are highlighted by orange lines. The distance between two INA rings belonging to two parallel equal layers (along b) or to two alternate neighboring layers (along c) is denoted by “INA ring parallel”, and “INA ring alternate”, respectively.

chlorine distance (HO(H)···Cl), and two oxygen−oxygen distances (H2O···(H)O; HO(H)···O) are considered (both the water molecule and the carboxyl moieties give rise to H bonds by using both their H and O atoms; see Figure S2 in the Supporting Information). Moreover, two intermolecular parameters, which represent the INA arene-stacking distances depicted in Figure 1, are reported in Tables 1 and 2. Both the “parallel” and the “alternate” configurations have been measured (INA ring parallel and INA ring alternate). If it is considered that long distances (3−4 to 7−9 Å) are involved, these parameters may give an idea of the structural impact of long-range dispersive effects. It can be observed from the data given in Table 1 that the Cl···O distance of 1α is well described, whereas major errors affect the related C−O−Cl angle (about 1.5°), thus resulting in uncertainty about the position of the −OH groups involved in the H bond. Unexpectedly, the parameters involving the Ru(II) atom are predicted with acceptable errors for both crystal forms at all levels of approximation used in the minimization of the crystal forms (see Tables S1 and S2 of the Supporting Information). 13 C and 1H NMR Calculations. The 13C MAS NMR spectra of the 1α and 1·H2O forms simulated by using the 13C isotropic chemical shifts computed on the B3LYP-D* optimized geometries are shown in Figure 2, together with their respective experimental spectra.2 The corresponding labeling scheme adopted is shown in Figure 3. A deeper comparison of the results obtained by means of the structures minimized with different functionals is reported in the Supporting Information. Overall, a satisfactory agreement is found, notwithstanding the perceivable peak shifts at 150−160 and ∼138 ppm in the 13 C MAS NMR theoretical spectrum of the 1α structure and the marked peak splitting at ∼120, ∼100, and ∼80 ppm in the 13 C MAS NMR spectrum of the 1·H2O form, not detectable by the experiment (Figure 2). This can be ascribed to the omission of thermal effects (calculations refer to 0 K) which are intrinsically present in the experiments.

Table 1. Cell Parameters and Important Angles and Distances of the 1α Polymorph Obtained from Full Geometry Optimizationsa a (Å) b (Å) c (Å) β (deg) V (Å3) Cl···O (Å) C−O−Cl (deg) N−Ru−p-cymene (deg) Ru−p-cymene (Å) Cl−Ru−Cl (deg) INA ring parallel (Å) INA ring alternate (Å)

B3LYP-D*

exptlb

14.890 (−0.867) 7.200 (−1.013) 15.123 (−2.433) 95.726 (0.287) 1613.20 (−4.304) 3.037 (0.004) 112.435 (1.538) 127.166 (0.188) 1.685 (0.016) 86.770 (0.629) 7.200 (0.074) 5.030 (0.075)

15.020 7.274 15.500 95.452 1685.75 3.033 113.973 127.354 1.669 86.141 7.274 5.105

a

Percent relative deviations are reported in parentheses for the cell parameters, and absolute deviations are reported in parentheses for the distances and angles. bFrom ref 2.

Overall, the cell parameters of the 1α polymorph are very well reproduced, the absolute errors being of less than 2.4%. The results obtained for the 1·H2O form can be considered satisfactory, although the angle β presents considerable deviations, and the volume of the unit cell for the optimized structure is smaller than the experimental value. This is probably due to a wrong inclusion of the long-range tail of the electrostatic part of H-bond interactions as dispersive interactions. Some relevant 1α polymorph distances and angles are also given in Table 1, together with their absolute deviation from the experimental data. The −Cl···(H)O distance (from now on referred as Cl···O) and the C−O(−H)···Cl− angle (from now on referred as C−O−Cl) are used to characterize the hydrogen bond length and angle of the 1α phase, respectively (Figure 1). In the same way, for 1·H2O (Table 2), the H2O oxygen− 7928

dx.doi.org/10.1021/ic5006743 | Inorg. Chem. 2014, 53, 7926−7935

Inorganic Chemistry

Article

Figure 3. Molecular structure and labeling scheme of [(p-cymene)Ru(κN-INA)Cl2]. The original NMR labeling from ref 2 for the carbon and oxygen atoms of the 1α polymorph is shown. The 1·H2O and 1β polymorphs and the two molecules of the 1γ polymorphs are labeled accordingly, independent of the mutual position of the −COOH and isopropyl groups (Figure 4).

previous works.2,6 The recalculated MADs are reported in parentheses in Table 3 and result in a value of 1.6 ppm. For the 1H NMR isotropic chemical shifts given in Table 4, the MADs computed for H1 and H5 δiso data are almost invariant. However, a marked deviation of about 7 ppm of the signal of H (OH), which corresponds to the hydrogen atom belonging to the carboxyl group in 1·H2O, should be highlighted. This discrepancy is probably due to a possible proton exchange with the water molecules that can occur during experiments and that is not taken into proper account by means of static calculations. A deeper investigation of the proton exchange and of its effects on the 1H chemical shifts should be based on molecular dynamics simulations that intrinsically include thermal effects.50−52 An extensive comparison of the results obtained by means of different levels of theory is reported in Tables S3 and S4 of Supporting Information. 1β Polymorph. Fully Optimized Structure. The 1β polymorph (P21/c, 164 atoms per cell) is characterized by two layers containing an intermolecular −COOH···Cl interaction, which lies in the ab plane (Figure S5, Supporting Information). Though its structure is similar to that of 1·H2O, the 1β phase presents the same kind of H bond (−COOH··· Cl) that belongs to the 1α polymorph. Cell Intra- and Intermolecular Parameters. The cell parameters obtained by B3LYP-D* full geometry optimizations and their percent relative deviations from experimental data are given in Table 5. The final cell is more compact than the experimental one, since the overestimation of the b and β parameters compensates for the symmetry change due to a and c. The H-bond pattern along a prevails over the b direction. In the same way, the superposition of aromatic groups provides a strong directionality to long-range correlation effects, consequently giving an underestimation of c. The results for the inter- and intramolecular parameters, also given in Table 5, underline a slight change of the N−Ru−pcymene and C−O−Cl angles (0.7 and 0.8°, respectively), whereas the Cl−Ru−Cl angle is wider than the experimental

Figure 2. Simulated and experimental 13C solid-state NMR spectra of the 1α, 1·H2O, 1β, and 1γ forms.

Further insights can be obtained by the analysis of the mean absolute deviations (MAD) between the computed and experimental 13C chemical shifts (δiso) data values of the 1α and 1·H2O crystal forms. MADs are computed as (∑ n |(δ iso,n (computed) − δiso,n(experimental)|)/n and are given in Table 3 for the B3LYP-D* optimized geometries. Overall, a good agreement is detected, the MAD being 1.3 ppm for 1α and 1.9 ppm for 1· H2O. A slightly larger deviation is expected for the 1·H2O chemical shifts, since the δiso scale was set by using the 1α structure. However, it is worth stressing that a major difference is noted between the computed and experimental 13C δiso assignments of the C1 and C5 atoms of 1·H2O. This implies a rotation of the whole INA moiety (C2 and C4 δiso values are experimentally equivalent). In fact, as is clearly shown in Figure 4a,b, the mutual positions of the −COOH and p-cymene isopropyl groups in the 1·H2O phase (C4−C3−C6−O1(H) −169.20°) are almost opposite with respect to those of the 1α polymorph (C4−C3−C6−O1(H) dihedral angle being 17.92°). Figure 4 shows the fully optimized structures; however, this situation is present also in the experimental nonrefined structures, but it has not been highlighted in 7929

dx.doi.org/10.1021/ic5006743 | Inorg. Chem. 2014, 53, 7926−7935

Inorganic Chemistry

Article

Table 3. Computed (Fully Optimized Structures) and Experimental 13C Chemical Shifts of the Four Crystalline Forms of [(pcymene)Ru(κN-INA)Cl2] δiso (ppm) 1α form B3LYP-D* C1 C2 C3 C4 C5 C6 C7 C8 C9 C10 C11 C12 C13 C14 C15 C16 MADc

155.4 126.8 136.6 123.1 157.0 165.8 101.1 88.7 87.7 97.8 86.7 79.0 16.8 33.5 24.4 16.0 1.3

1·H2O form exptla 156.9 125.8 139.7 122.8 158.4 164.6 101.2 86.9 86.9 96.0 86.3 77.9 17.7 31.7 25.0 18.6

B3LYP-D* 152.4 124.5 134.2 122.7 157.0 166.7 99.4 93.9 88.5 102.5 83.1 79.6 19.9 33.6 25.1 17.6 1.9(1.6)d

1β form

1γ form

exptla

B3LYP-D*

exptla

B3LYP-D*

exptlb

d

151.1 123.9 138.7 125.2 159.6 165.7 102.2 90.7 90.7 100.1 84.8 78.7 19.6 34.0 23.6 23.0 3.3(1.8)d

152.9 122.2 140.1 124.8 161.4 164.5 103.1 77.6d 82.4d 97.1 87.8d 87.3d 19.4 31.8 21.9 26.5

151.4/153.5 122.4/125.1 139.4/132.5 124.4/119.8 158.5/160.3 165.2/162.7 101.6/101.9 92.2/91.6 89.3/90.3 102.2/101.5 81.1/84.3 81.7/80.5 20.9/19.3 31.8/32.7 23.9/24.4 18.8/21.2 3.1(2.3)d

160.3/159.5d 120.8/120.0d 139.7/137.1 123.3/122.2d 153.8/153.3d 167.3/162.9 100.5/99.9 79.2 89.9/87.0 97.4 86.5/84.3 83.2/80.5 19.2/17.3 30.8 26.8/24.9 19.2

159.6 122.6 137.7 122.6 154.5d 167.2 100.1 92.1 86.1 100.1 80.5 80.5 20.2 31.7 26.3 18.4

a From ref 2. bFrom ref 6. cMean absolute deviations (reported in boldface). For 1γ, MADs are calculated on the mean of δiso data values of molecules (1) and (2). dThe exchange of C1/C5 δiso values for 1·H2O, C8/C12 and C9/C11 δiso values for 1β, and C1/C5 and C2/C4 δiso values for 1γ (reported in boldface) yields the MADs reported in parentheses.

Information), which is comparable to the absolute percent relative deviations of the Cl−Ru−Cl angle. 13 C and 1H NMR Calculations. Figure 2 allows the comparison of the 13C MAS NMR experimental spectrum of the 1β polymorph with the computed spectrum. Moreover, the computed 13C and 1H chemical shifts (δiso) are given in Tables 3 and 4, together with the experimental values.2 The impact of geometry optimization of the XRPD 1β structure on the agreement between theoretical and experimental spectra is notable. Full details are available in the Supporting Information. Briefly, the agreement between the δiso data values calculated for the nonrefined XRPD experimental structure and the experimental structure is poor. The optimization of H atoms ameliorates significantly the 13C and 1 H δiso data values and the 13C spectrum with respect to the nonrefined values, while only minor differences are observed when the hydrogen-optimized and fully optimized spectra are compared. Importantly, the computed δiso data values of the aromatic −CH− groups (Tables 3 and 4) seem to be in better agreement with the experiment if the δiso assignments of the C9 and C8 carbon atoms (which belong to the INA p-cymene ring) are interchanged with those of C11 and C12, respectively. In fact, the MAD values for the 13C/1H NMR signals after the 8/12 and 9/11 atom exchange are MAD(13C) = 1.8 ppm MAD(1H) = 0.5 ppm. These figures underline a significant improvement with respect to the assignment by Chierotti et al.,6 which yields MAD(13C) = 3.3 ppm and MAD(1H) = 0.6 ppm (see Tables 3 and 4). To support this hypothesis, it is worth recalling that Chierotti et al.6 detected uncertainty in the aromatic hydrogens and labeled the assignments of 1H δiso as H9/H11, H11/H9, H8/ H12, and H12/H8. However, in the present work the original atom labeling from ref 2 has been retained (Figure 3). 1γ Polymorph. Fully Optimized Structure. The phase 1γ belongs to the P21/a space group, a nonstandard setting of P21/

Figure 4. (top) Comparison between the −COOH group structural positions of the different polymorphs (fully optimized geometries): (a) 1α; (b) 1·H2O; (c) 1β; (d, e) the two molecules (1) and (2), respectively, of 1γ. (bottom) The two H bonds in the 1γ form (fully optimized geometry), highlighted with dotted lines.

value of about 7°. The Cl−Ru−Cl moiety lies perpendicular to the c direction, along which the interplay of the H-bond network and vdW forces result in an overestimation of the cell packing. Although the errors in all the distance parameters are small and fall within ∼0.3 Å, it is worth noting the large absolute percent relative deviations of the Cl···O distance (involving the hydrogen bond) displayed in Figure S6 (Supporting 7930

dx.doi.org/10.1021/ic5006743 | Inorg. Chem. 2014, 53, 7926−7935

Inorganic Chemistry

Article

Table 4. Computed (Fully Optimized Structures) and Experimental 1H Chemical Shifts of the Four Crystalline Forms of [(pcymene)Ru(κN-INA)Cl2] δiso (ppm) 1α form B3LYP-D* H1 H2 H4 H5 H(OH) H8 H9 H11 H12 H13e H14 H15e H16e MADc

8.8 8.8 6.4 7.8 10.2 4.8 4.2 4.9 4.7 2.7 3.2 1.5 0.6 0.3

1·H2O form exptla 9.4 8.7 6.2 8.6 9.9b 4.6 4.6 5.0 4.8 2.5 3.2 1.2 0.5

B3LYP-D* 8.0 7.2 6.4 8.3 15.6 5.7 5.4 4.5 5.2 2.3 3.3 2.0 1.7 0.9(0.9)d

1β form exptla d

8.6 6.3d 6.7d 8.9d 8.7 5.8 5.9 5.1 5.1 1.8 3.1 1.4 1.3

1γ form

B3LYP-D*

exptla

B3LYP-D*

exptlb

7.7 7.0 7.2 8.4 10.8 5.4 5.5 4.6 5.5 2.5 3.3 2.3 1.3 0.6(0.5)d

8.5 7.1

8.2/8.3 7.1/7.1 7.1/7.4 8.4/8.9 11.0/10.6 5.6/5.9 5.7/5.5 4.7/4.8 5.1/5.3 2.6/2.6 4.2/3.4 2.3/2.3 1.6/1.2 0.5(0.5)d

9.1d 7.1d 6.6/6.5d 8.8/8.9d 10.5 5.0 5.5 5.6/5.4 5.0/5.4 1.7/1.4 3.2 1.8/1.6 1.4

9.1 10.1 5.3d 4.8d 6.0d 5.5d 1.8 3.2 1.2 1.6

a From ref 2. bFrom ref 6. cMean absolute deviations (reported in boldface). For 1γ, MADs are calculated on the mean of δiso data values of molecules (1) and (2). dThe exchange of H1/H5 and H2/H4 δiso values for 1·H2O, H8/H12 and H9/H11 δiso values for 1β, and H1/H5 and H2/ H4 δiso values for 1γ (reported in boldface) yields the MADs reported in parentheses. eOnly one experimental δiso data value is given for the three H atoms of CH3 groups (H13, H15, and H16); the mean value is reported for the computed data.

The first molecule in the fully optimized structure of the 1γ polymorph, (also referred as (1) in the following) (Figure 4d), being characterized by the −COOH dihedral angle of 150.21° skewed with respect to the INA ring, is similar to the 1β form (Figure 4c). The second molecule (also referred as (2) in the following) (Figure 4e) presents a dihedral angle of 1.15° (almost planar) which is comparable to that of 1α (Figure 4a), although the acidic hydrogen points in a different direction with respect to the 1α polymorph. In both molecules of the 1γ polymorph, the acidic hydrogen establishes an H···Cl bond with similar geometry, as highlighted in Figure 5: the O···Cl distances are 3.080 and 3.187 Å for molecules (1) and (2), respectively, and the O−H···Cl angles are 161.88 and 162.45° for molecules (1) and (2), respectively. This is reflected also by the electron charge density of the H and Cl atoms involved in the COOH···Cl interaction. A qualitative pictorial analysis for the three fully optimized (B3LYP-D*) anhydrous polymorphs is given in Figure 6. The Bader atomic volumes of the charge density around H and Cl atoms, partitioned according to Bader’s quantum theory of atom in molecules (QTAIM),53 are depicted with an isosurface density value of 0.02 a0−3. The figure shows that the partial charge on H atoms tends to delocalize in the direction of acceptors (Cl atoms) and the charge densities for the H···Cl interaction are similarly shaped for the four molecules, supporting the hypothesis that the two H bonds present in the 1γ phase have similar strengths. Indeed, both the nonrefined XRPD experimental structure (Figure 5) and the results of the refined XRPD/NMR/DFT structural characterization6 furnish a detectable difference in the geometry of the O−H···Cl moiety. In fact, the O···Cl distances are 3.103/3.170 Å (Figure 5) and 3.085/3.183 Å, and the O− H···Cl angles are 162.65/151.48° (Figure 5) and 156.2/166.2° for the molecules (1) and (2) of the nonrefined XRPD experimental and refined XRPD/NMR/DFT structures,6 respectively. This evidence led Chierotti et al.6 to hypothesize the presence of one strong H···Cl bond and a weaker H···Cl contact in the 1γ polymorph. Nonetheless, the structure and

Table 5. Cell Parameters and Important Angles and Distances of the 1β Polymorph Obtained from Full Geometry Optimizationsa a (Å) b (Å) c (Å) β (deg) V (Å3) Cl···O (Å) C−O−Cl (deg) N−Ru−p-cymene (deg) Ru−p-cymene (Å) Cl−Ru−Cl (deg) INA ring parallel (Å) INA ring alternate (Å)

B3LYP-D*

exptlb

14.103 (−4.092) 15.444 (2.375) 7.410 (−4.054) 99.936 (1.928) 1589.91 (−6.28) 3.130 (0.294) 124.006 (0.712) 129.081 (0.846) 1.687 (0.019) 87.943 (7.395) 7.410 (0.313) 3.792 (0.158)

14.705 15.086 7.723 98.046 1696.54 3.424 123.294 128.235 1.668 80.548 7.723 3.950

a

Percent relative deviations are reported in parentheses for the cell parameters, and absolute deviations are reported in parentheses for the distances and angles. bFrom ref 6.

c for which the polymorph contains two molecules of [(pcymene)Ru(κN-INA)Cl2] in the asymmetric unit (8 molecules/328 atoms per cell). Substantial differences in the spatial position of the −COOH group in the two molecular units are observed. A description of the fully optimized structures is given in Figure 5, where a comparison with the nonrefined XRPD experimental structure, used as input, is also reported. A comparison among the four crystal forms furnishes interesting information (Figure 4): the −COOH moiety of the 1α polymorph is almost planar with respect to the INA ring, the C4−C3−C6−O1(H) dihedral angle being 17.92° (Figure 4a). As explained in the previous section, in the 1·H2O phase a substantial rotation of the −COOH group is observed (C4−C3−C6−O1(H) −169.20°; Figure 4b), whereas the C4− C3−C6−O1(H) dihedral angle of the 1β polymorph is 143.29° (Figure 4c). 7931

dx.doi.org/10.1021/ic5006743 | Inorg. Chem. 2014, 53, 7926−7935

Inorganic Chemistry

Article

Figure 5. ab plane view of (a) the fully optimized structure (B3LYP-D*) of the 1γ form and (b) the nonrefined XRPD experimental structure used as input for calculations.

Table 6. Cell Parameters and Important Angles and Distances of the 1γ Polymorph Obtained from Full Geometry Optimizationsa a (Å) b (Å) c (Å) β (deg) V (Å3) Cl···O (Å) (1) (2) C−O−Cl (deg) (1) (2) N−Ru−p-cymene (deg) (1) (2) Ru−p-cymene (Å) (1) (2) Cl−Ru−Cl (deg) (1) (2) INA ring parallel (Å) INA ring alternate (Å)

Figure 6. Atomic volume partitioning of the Bader charge around the atoms involved in the COOH···Cl interaction for the (a) 1α, (b) 1β, (c) 1γ(1), and (d) 1γ(2) crystal forms. The yellow surface around the H and Cl atoms is represented with an isosurface density value of 0.02 a0−3. The original charge density has been partitioned through the software “Bader”.54−56

labeling of molecule (2) in the asymmetric unit was not reported in ref 6. Cell Inter- and Intramolecular Parameters. The full geometry optimization leads to minor changes in cell parameters. The direction a, along which both H bonds and dispersive interactions are present, constitutes an exception, being underestimated by 2.4 Å (Table 6). As for 1β, the INA rings of alternate layers are aligned along c. Marked absolute percent relative deviations on intermolecular parameters are related to the C−O−Cl angle (C−O−Cl 3.6° for (1) and 9.9° for (2)), which involves the hydrogen bonds (Figure S8, Supporting Information). Moreover, a considerable discrepancy with the experimental structure is

B3LYP-D*

exptlb

28.557 (−7.630) 15.613 (1.773) 7.309 (−0.791) 95.980 (0.740) 3241.32 (−6.85)

30.916 15.341 7.368 95.275 3479.64

3.080 (0.023) 3.187 (0.017)

3.103 3.170

124.802 (3.604) 130.037 (9.957)

128.406 139.994

128.440 (2.388) 128.286 (4.337)

126.052 132.623

1.685 (0.017) 1.690 (0.022)

1.668 1.668

88.756 (3.213) 87.368 (3.708) 7.309 (0.059) 3.798 (0.037)

85.543 83.660 7.368 3.835

a

Percent relative deviations are reported in parentheses for the cell parameters, and absolute deviations are reported in parentheses for the distances and angles. bFrom ref 6.

7932

dx.doi.org/10.1021/ic5006743 | Inorg. Chem. 2014, 53, 7926−7935

Inorganic Chemistry

Article

These results reflect the differences in the accuracy of the two theoretical approaches and suggest that the peaks of both hydrogen atoms engaged in the formation of H bonds lie in the same spectral regions. That is, the peak of the second hydrogen does not fall overlapped under the signals of aromatic hydrogen atoms in ortho positions with respect to the pyridine nitrogen atoms, as hypothesized by Chierotti et al.6 The NMR-GIPAW DFT calculations carried out in this work on the structures optimized by including the dispersion corrections furnish results more comparable with the experimental data, as testified to by the impressive agreement with the 1H−13C FSLG-HETCOR spectrum shown in Figure 7.

observed for the optimized Cl−Ru−Cl intramolecular angle (overestimated by ∼3°). Such rearrangement competes with ring-stacking dispersive interactions to the mild shortening of the c direction. Substantial differences are observed also for the N−Ru−p-cymene angle (+2.4°/−4.3° for molecules (1) and (2), respectively). 13 C and 1H NMR Calculations. The simulated 13C MAS NMR spectra of the 1γ polymorph computed on the refined structure is shown in Figure 2, and the calculated 13C and 1H chemical shifts (δiso) are given in Tables 3 and 4. An extensive comparison of the results obtained by means of the structures minimized with different functionals is reported in the Supporting Information. The presence of two molecules of [(p-cymene)Ru(κNINA)Cl2] in the asymmetric unit, almost indistinguishable in the experimental spectra, yields two different sets of δiso data values for the calculated parameters and a manifold 13C spectrum, where many peaks are overlapped. Therefore, the agreement between the experimental and simulated spectra can be considered only qualitative. Importantly, more accuracy in the peak position of the carbon atoms belonging to the ring carrying the −COOH group of the two molecules (Table 3) is achieved when the δiso data value assignment of the aromatic C1 is exchanged with its equivalent atom C5, and C2 with C4. This is highlighted by the values of the MADs calculated by considering the mean δiso data values of molecules (1) and (2), both with the original assignment and after the exchange. A comparison between the computed and experimental 1H δiso data values (Table 4) is difficult, since the broad experimental signals impede the discrimination and assignments of most of the 1H resonances for the symmetrynonequivalent molecules. However, in general, the same considerations made for the 13C δiso data values hold also in this case. Interestingly, a broad signal centered at 10.5 ppm was detected for the −COOH hydrogen (H6) in the experimental MAS spectrum recorded by Chierotti et al.6 This signal presents a reduced intensity with respect to the corresponding signal in the 1β polymorph. Thus, the authors hypothesized the presence of a second weak hydrogen bond, whose signal is supposed to fall in the region characteristic of the pyridine protons. This hypothesis was supported by the results of a NMR-GIAO DFT calculation, which yielded 1H δiso data values of 12.2 and 9.6 ppm for the strong and weak interactions, respectively. However, no signal is present at 12.2 ppm in the experimental spectrum. Moreover, the more accurate NMRGIPAW DFT calculations carried out in this work on the nonrefined experimental geometry (Figure S9, Supporting Information) yields a very different result: 1H δiso values of 9.5 and 9.4 ppm for the acidic hydrogen of the two molecules. These discrepancies suggest that, as pointed out previously, the experimental geometry might require further refinement. In fact, it is worth recalling that the experimental structure was obtained by an iterative XRPD-DFT calculation procedure and no correction for dispersion effects was included in the periodic lattice DFT calculations.6 Moreover, NMR-GIPAW DFT calculations on the refined XRPD/NMR/DFT structure reported by Chierotti et al.6 provide 1H δiso data values of 11.0 and 10.9 ppm for the strong and weak H bonds and are in disagreement with the DFTGIAO results.

Figure 7. Superposition of the 1H−13C FSLG-HETCOR spectrum obtained for the 1γ crystal form (in black; ref 6) and the correlated 1 H−13C NMR parameters (blue dots) obtained by means of the NMR-GIPAW DFT calculations on the B3LYP-D* fully optimized structure.

In particular, the 1H δiso data values computed for the −COOH hydrogen (H6) after hydrogen relaxation (Table S6, Supporting Information) are 10.4/10.6 ppm for molecules (1)/(2), respectively. Slightly different values (11.0/10.6 ppm for molecules (1)/(2), respectively) are obtained also in the case of the fully relaxed structure (Table 4 and Table S6), once again endorsing the hypothesis of two distinct COOH···Cl bonds of comparable strength.



CONCLUSIONS The 13C/1H NMR parameters and 13C MAS NMR spectra of the 1α, 1·H2O, 1β, and 1γ forms of [(p-cymene)Ru(κNINA)Cl2] have been calculated by means of the PBE-GIPAW method. The results showed that, in order to obtain good agreement with the spectroscopic data, full geometry optimizations of the experimental structures or, at least, computational refinement of the hydrogen atoms is mandatory. Moreover, addition of a dispersion correction term (the “classical” Grimme dispersion correction −D2 or its alternative reparametrization devised for crystals −D*) to the pure B3LYP functional used, under periodic boundary conditions, for the geometry optimization protocol is extremely important to ameliorate the structure and spectra prediction. The experimental and computed 13C and 1H δiso data value differences, expressed as mean absolute deviations, lie in the ranges 1.3−2.9 and 0.3−1.0 ppm, respectively. Furthermore, some revisions in the experimental assignment of the 13C/1H NMR δiso parameters of the 1·H2O, 1β, and 1γ forms have been suggested. The mismatch in the assignment seems to be mainly 7933

dx.doi.org/10.1021/ic5006743 | Inorg. Chem. 2014, 53, 7926−7935

Inorganic Chemistry

Article

due to the rotation of the −COOH moiety, which occurs at the 1α−1·H2O transition and was not considered in the experiments. Finally, the results obtained suggest the presence of two COOH···Cl hydrogen bonds of comparable strength established by the two molecules in the asymmetric unit of the 1γ polymorph, in partial disagreement with previous findings.



(12) Pedone, A.; Gambuzzi, E.; Menziani, M. C. J. Phys. Chem. C 2012, 116, 14599−14609. (13) Pedone, A.; Charpentier, T.; Menziani, M. C. J. Mater. Chem. 2012, 22, 12599−12608. (14) Charpentier, T.; Menziani, M. C.; Pedone, A. RSC Adv. 2013, 3, 10550−10578. (15) Harris, R. K.; Ghi, P. Y.; Puschmann, H.; Apperley, D. C.; Griesser, U. J.; Hammond, R. B.; Ma, C.; Roberts, K. J.; Pearce, G. J.; Yates, J. R.; Pickard, C. J. Org. Process Res. Dev. 2005, 9, 902−910. (16) Harris, R. K.; Hodgkinson, P.; Larsson, T.; Muruganantham, A.; Ymén, I.; Yufit, D. S.; Zorin, V. Cryst. Growth Des. 2008, 8, 80−90. (17) Pickard, C. J.; Mauri, F. Phys. Rev. B 2001, 63, 245101. (18) Yates, J. R.; Pickard, C. J.; Mauri, F. Phys. Rev. B 2007, 76, 024401. (19) Brouwer, D. H.; Enright, G. D. J. Am. Chem. Soc. 2008, 130, 3095−3105. (20) Brouwer, D. H. J. Am. Chem. Soc. 2008, 130, 6306−6307. (21) Day, G. M.; Motherwell, W. D. S.; Ammon, H. L.; Boerrigter, S. X. M.; Della Valle, R. G.; Venuti, E.; Dzyabchenko, A.; Dunitz, J. D.; Schweizer, B.; van Eijck, B. P.; Erk, P.; Facelli, J. C.; Bazterra, V. E.; Ferraro, M. B.; Hofmann, D. W. M.; Leusen, F. J. J.; Liang, C.; Pantelides, C. C.; Karamertzanis, P. G.; Price, S. L.; Lewis, T. C.; Nowell, H.; Torrisi, A.; Scheraga, H. A.; Arnautova, Y. A.; Schmidt, M. U.; Verwer, P. Acta Crystallogr., Sect. B 2005, 61, 511−527. (22) Day, G. M.; Cooper, T. G.; Cruz-Cabeza, A. J.; Hejczyk, K. E.; Ammon, H. L.; Boerrigter, S. X. M.; Tan, J. S.; Della Valle, R. G.; Venuti, E.; Jose, J.; Gadre, S. R.; Desiraju, G. R.; Thakur, T. S.; van Eijck, B. P.; Facelli, J. C.; Bazterra, V. E.; Ferraro, M. B.; Hofmann, D. W. M.; Neumann, M. A.; Leusen, F. J. J.; Kendrick, J.; Price, S. L.; Misquitta, A. J.; Karamertzanis, P. G.; Welch, G. W. A.; Scheraga, H. A.; Arnautova, Y. A.; Schmidt, M. U.; van de Streek, J.; Wolf, A. K.; Schweizer, B. Acta Crystallogr., Sect. B 2009, 65, 107−125. (23) Bardwell, D. A.; Adjiman, C. S.; Arnautova, Y. A.; Bartashevich, E.; Boerrigter, S. X. M.; Braun, D. E.; Cruz-Cabeza, A. J.; Day, G. M.; Della Valle, R. G.; Desiraju, G. R.; van Eijck, B. P.; Facelli, J. C.; Ferraro, M. B.; Grillo, D.; Habgood, M.; Hofmann, D. W. M.; Hofmann, F.; Jose, K. V. J.; Karamertzanis, P. G.; Kazantsev, A. V.; Kendrick, J.; Kuleshova, L. N.; Leusen, F. J. J.; Maleev, A. V.; Misquitta, A. J.; Mohamed, S.; Needs, R. J.; Neumann, M. A.; Nikylov, D.; Orendt, A. M.; Pal, R.; Pantelides, C. C.; Pickard, C. J.; Price, L. S.; Price, S. L.; Scheraga, H. A.; van de Streek, J.; Thakur, T. S.; Tiwari, S.; Venuti, E.; Zhitkov, I. K. Acta Crystallogr., Sect. B 2011, 67, 535−551. (24) Grimme, S. J. Comput. Chem. 2006, 27, 1787−1799. (25) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. J. Chem. Phys. 2010, 132, 154104. (26) Becke, A. D.; Johnson, E. R. J. Chem. Phys. 2005, 122, 154104. (27) Johnson, E. R.; Becke, A. D. J. Chem. Phys. 2005, 123, 024101. (28) Becke, A. D.; Johnson, E. R. J. Chem. Phys. 2006, 124, 014104. (29) Becke, A. D.; Johnson, E. R. J. Chem. Phys. 2007, 127, 154108. (30) Tkatchenko, A.; Scheffler, M. Phys. Rev. Lett. 2009, 102, 073005. (31) Dion, M.; Rydberg, H.; Schröder, E.; Langreth, D. C.; Lundqvist, B. I. Phys. Rev. Lett. 2004, 92, 246401. (32) Vydrov, O. A.; Van Voorhis, T. J. Chem. Phys. 2010, 133, 244103. (33) Schatschneider, B.; Liang, J.; Reilly, A. M.; Marom, N.; Zhang, G. X.; Tkatchenko, A. Phys. Rev. B 2013, 87, 060104. (34) Pedone, A.; Presti, D.; Menziani, M. C. Chem. Phys. Lett. 2012, 541, 12−15. (35) Presti, D.; Pedone, A.; Menziani, M. C.; Civalleri, B.; Maschio, L. CrystEngComm 2014, 16, 102−109. (36) Dovesi, R.; Saunders, V. R.; Roetti, C.; Orlando, R.; ZicovichWilson, C. M.; Pascale, F.; Civalleri, B.; Doll, K.; Harrison, N. M.; Bush, I. J.; D’Arco, P.; Llunell, M. CRYSTAL09 User’s Manual; University of Torino: Torino, Italy, 2009. (37) Dovesi, R.; Orlando, R.; Civalleri, B.; Roetti, C.; Saunders, V. R.; Zicovich-Wilson, C. M. Z. Kristallogr. 2005, 220, 571−573. (38) Becke, A. D. J. Chem. Phys. 1993, 98, 5648−5652. (39) Stephens, P.; Devlin, F.; Chabalowski, C.; Frisch, M. J. J. Phys. Chem. 1994, 98, 11623−11627.

ASSOCIATED CONTENT

S Supporting Information *

Text, tables, and figures giving the linear regression for 13C/1H NMR δiso of 1α as well as the descriptions and figures of the four crystalline structures, all of the computed 13C/1H NMR δiso values with the related 13C spectra, and the calculated 13 C/1H NMR correlation spectra. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Authors

*E-mail for A.P.: [email protected]. *E-mail for M.C.M.: [email protected]. Author Contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors acknowledge Prof. Alessia Bacchi for the nonrefined structures of 1β and 1γ and Dr. Ilaria Ciofini for useful discussions. D.P. thanks Regione Emilia-Romagna for financial support through the SPINNER 2013 program entitled “Ottimizzazione delle forme molecolari e crystalline di farmaci, fitofarmaci, pesticidi in relazione ad attività, biodisponibilità aspetti brevettuali, e alla produzione di polimorfi, solvati e cocrystalli con metodi a basso impatto ambientale”.



REFERENCES

(1) Soldatov, D. V. J. Chem. Crystallogr. 2006, 36, 747−768. (2) Bacchi, A.; Cantoni, G.; Chierotti, M. R.; Girlando, A.; Gobetto, R.; Lapadula, G.; Pelagatti, P.; Sironi, A.; Zecchini, M. CrystEngComm 2011, 13, 4365−4375. (3) Bacchi, A.; Bosetti, E.; Carcelli, M.; Pelagatti, P.; Rogolino, D. Eur. J. Inorg. Chem. 2004, 10, 1985−1991. (4) Bacchi, A.; Bosetti, E.; Carcelli, M.; Pelagatti, P.; Rogolino, D. CrystEngComm 2004, 6, 177−183. (5) Bacchi, A.; Bosetti, E.; Carcelli, M.; Pelagatti, P.; Rogolino, D.; Pelizzi, G. Inorg. Chem. 2005, 44, 431−442. (6) Chierotti, M. R.; Gobetto, R.; Nervi, C.; Bacchi, A.; Pelagatti, P.; Colombo, V.; Sironi, A. Inorg. Chem. 2014, 53, 139−146. (7) Santos, S. M.; Rocha, J.; Mafra, L. Cryst. Growth Des. 2013, 13, 2390−2395. (8) Többens, D. M.; Glinneman, J.; Chierotti, M. R.; van de Streek, J.; Sheptyakov, D. CrystEngComm 2012, 14, 3046−3055. (9) Bonhomme, C.; Gervais, C.; Babonneau, F.; Coelho, C.; Pourpoint, F.; Azaïs, T.; Ashbrook, S. E.; Griffin, J. M.; Yates, J. R.; Mauri, F.; Pickard, C. J. Chem. Rev. 2012, 112, 5733−5779. (10) Ashbrook, S. E.; Berry, A. J.; Frost, D. J.; Gregorovic, A.; Pickard, C. J.; Readman, J. E.; Wimperis, S. J. Am. Chem. Soc. 2007, 129, 13213−13224. (11) Hinklin, T.; Toury, B.; Gervais, C.; Babonneau, F.; Gislason, J.; Morton, R.; Laine, R. Chem. Mater. 2004, 16, 21−30. 7934

dx.doi.org/10.1021/ic5006743 | Inorg. Chem. 2014, 53, 7926−7935

Inorganic Chemistry

Article

(40) Civalleri, B.; Zicovich-Wilson, C. M.; Valenzano, L.; Ugliengo, P. CrystEngComm 2008, 10, 405−410. (41) Peterson, K.; Figgen, D.; Dolg, M.; Stoll, H. J. Chem. Phys. 2007, 126, 124101. (42) Igel-Mann, G.; Stoll, H.; Preuss, H. Mol. Phys. 1988, 65, 1321. (43) http://www.theochem.uni-stuttgart.de/pseudopotentials/. (44) Pickard, C.; Mauri, F. Phys. Rev. B 2001, 63, 245101. (45) Yates, J.; Pickard, C.; Mauri, F. Phys. Rev. B 2007, 76, 024401. (46) Segall, M. D.; Lindan, P. J. D.; Probert, M. J.; Pickard, C. J.; Hasnip, P. J.; Clark, S. J.; Payne, M. C. J. Phys.: Condens. Matter 2002, 14, 2717. (47) Clark, S.; Segall, M.; Pickard, C.; Hasnip, P.; Probert, M.; Refson, K.; Payne, M. Z. Kristallogr. 2005, 220, 567−570. (48) Perdew, J.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77, 3865−3868. (49) Pedone, A.; Charpentier, T.; Menziani, M. C. Phys. Chem. Chem. Phys. 2010, 12, 6054−6066. (50) De Gortari, I.; Portella, G.; Salvatella, S.; Bajaj, V. S.; van der Wel, P. C. A.; Yates, J. R.; Segall, M. D.; Pickard, C. J.; Payne, M. C.; Vendruscolo, M. J. Am. Chem. Soc. 2010, 132, 5993−6000. (51) Dumez, J.-N.; Pickard, C. J. J. Chem. Phys. 2009, 130, 104701− 104708. (52) Barone, V.; Cimino, P.; Pedone, A. Magn. Reson. Chem. 2010, 48, S11−S22. (53) Bader, R. F. W. Chem. Rev. 1991, 130, 893−928. (54) Tang, W.; Sanville, E.; Henkelman, G. J. Phys.: Condens. Matter 2009, 21, 084204. (55) Sanville, E.; Kenny, S. D.; Smith, R.; Henkelman, G. J. Comput. Chem. 2007, 28, 899−908. (56) Henkelman, G.; Arnaldsson, A.; Jónsson, H. Comput. Mater. Sci. 2006, 36, 354−360.

7935

dx.doi.org/10.1021/ic5006743 | Inorg. Chem. 2014, 53, 7926−7935